Hostname: page-component-7c8c6479df-nwzlb Total loading time: 0 Render date: 2024-03-28T13:37:36.277Z Has data issue: false hasContentIssue false

Resolved Observations of Transition Disks

Published online by Cambridge University Press:  06 April 2016

Simon Casassus*
Affiliation:
Departamento de Astronomía, Universidad de Chile, Casilla 36-D, Santiago, Chile
Rights & Permissions [Opens in a new window]

Abstract

Resolved observations are bringing new constraints on the origin of radial gaps in protoplanetary disks. The kinematics, sampled in detail in one case-study, are indicative of non-Keplerian flows, corresponding to warped structures and accretion which may both play a role in the development of cavities. Disk asymmetries seen in the radio continuum are being interpreted in the context of dust segregation via aerodynamic trapping. We summarise recent observational progress, and describe prospects for improvements in the near term.

Type
Review Article
Copyright
Copyright © Astronomical Society of Australia 2016 

1 INTRODUCTION

Circumstellar disks evolution is an aspect of stellar formation (Shu, Adams, & Lizano Reference Shu, Adams and Lizano1987; Zuckerman Reference Zuckerman2001; Williams & Cieza Reference Williams and Cieza2011; Dunham et al. Reference Dunham2015). In Class II young-stellar-objects, the dissipation of the protostellar envelope exposes a pre-main-sequence star surrounded by a gaseous accretion disk. In this framework, the spectral-energy-distributions (SEDs) provide constraints on radial disk structure. Central cavities (and gaps) are inferred in so-called transition disks (and pre-transitional disks, Espaillat, Calvet, D’Alessio, Hernández, Qi, Hartmann, Furl Reference Espaillat, Calvet, D’Alessio, Hernández, Qi, Hartmann, Furlan and Watson2007, TDs hereafter), whose structure is being refined with resolved observations.

This review provides a selection of structures seen in TDs, with an emphasis on the radio domain. Insights on the origin of the cavities based on TD demographics are summarised in the accompanying review by Owen (Reference Owen2015). We caution that resolved observations of TDs are affected by many biases, as they target the brightest sources with the largest cavities, so that they cannot be extrapolated to reflect the whole population (which is for now studied by complete photometric surveys, e.g. Cieza et al. Reference Cieza, Schreiber, Romero, Williams, Rebassa-Mansergas and Merín2012, Reference Cieza, Williams, Kourkchi, Andrews, Casassus, Graves and Schreiber2015). Instead, such resolved observations inform on the range of possible astrophysical phenomena at work in disks.

The observational identification of abrupt disk inclination changes (Section 2) provides the information on disk orientation required to interpret the dynamics of residual gas inside TD cavities (Section 3). The outer rings surrounding the central cavities of most TDs resolved so far in the submillimetre continuum show non-axial symmetry, with intensity contrasts from a few to ~ 100 (Section 4). Complex spiral patterns are seen in the outer disks, which are also beginning to be resolved in molecular lines (Section 5). The exciting observation of giant protoplanet candidates opens an enormous research potential, since their formation is expected to carve a radial gap, but their detection in TDs is still far from systematic (Section 6). The inferred warped accretion flows, the dramatic non-axial symmetry of some outer disks, and the observed spiral patterns, motivate questions on their origins, their connection to embedded protoplanets, and on their role in disk evolution (Section 7).

2 WARPS

Warps, or variable inclination and orientation as a function of stellocentric distance, have long been suspected to occur in circumstellar disks. In this Section, we compile evidence for warps in young circumstellar disks in a broader context, with special attention to TDs and the case of HD 142527.

2.1. Debris disks

Warps are seen in debris disks, or Class III young stellar objects that have dissipated the left-over primordial gas. This is the case of the edge-on warp seen in β Pic (Golimowski et al. Reference Golimowski2006; Millar-Blanchaer et al. Reference Millar-Blanchaer2015). Another example is AU Mic (Wang et al. Reference Wang2015; Boccaletti et al. Reference Boccaletti2015), and perhaps also HD 110058 (Kasper et al. Reference Kasper, Apai, Wagner and Robberto2015). The identification of warps in debris disks requires an edge-on view, and the lack of gaseous counterparts hampers reaching definitive conclusions on their structure, and so distinguish continuous warps from superimposed disks.

2.2. Indications of warps in gas-rich disks

Inner warps, or a tilted inclination at the disk centre, could lead to the photometric variations seen in some high-inclination (so close to edge-on) Class II classical T-Tauri starsFootnote 1 , such as in RY Lup (Manset et al. Reference Manset, Bastien, Ménard, Bertout, Le van Suu and Boivin2009), TWA 30 (Looper et al. Reference Looper2010), and AA Tau (Bouvier et al. Reference Bouvier, Grankin, Ellerbroek, Bouy and Barrado2013). The light curve variability of AA Tau is well accounted for by a magnetically induced warp on scales of a few stellar radii or ≲ 0.1 AU (Esau, Harries, & Bouvier Reference Esau, Harries and Bouvier2014), and part of the occulting structure could extend beyond 1 AU (Schneider et al. Reference Schneider, France, Günther, Herczeg, Robrade, Bouvier, McJunkin and Schmitt2015). In the NGC 2264 open cluster, up to ~ 40% of young and gas-rich cTTs with thick inner disks present AA Tau-like variability (Alencar et al. Reference Alencar2010), suggesting that such inner warps represent a fairly common phase in the early evolution of circumstellar disks (McGinnis et al. Reference McGinnis2015). Light curve variability on weeks and months timescales, due to obscuring material, has also been reported in the TDs T Cha (Schisano et al. Reference Schisano, Covino, Alcalá, Esposito and Guenther2009) and GM Aur (Ingleby et al. Reference Ingleby, Espaillat, Calvet, Sitko, Russell and Champney2015).

In TDs, warps have also been hinted at in the form of significant inclination changes in observations with different angular resolutions. For example in GM Aur, with a stellar mass ≲ 1 M, Hughes et al. (Reference Hughes2009) propose a central warp to explain the small change in disk position angleFootnote 2 of 11° ± 2° when comparing the major axis of the submillimetre continuum sampled at 0.3 arcsec, with that of the CO(3 − 2) kinematicsFootnote 3 sampled at 2 arcsec. Likewise, Tang et al. (Reference Tang, Guilloteau, Piétu, Dutrey, Ohashi and Ho2012) explain that in AB Aur (~ 2 M), the inclination inferred from near-IR interferometry (i ~ 20° within the central AU; Eisner et al. Reference Eisner, Lane, Hillenbrand, Akeson and Sargent2004) is close to that sampled over 20 AU scales, and lower than sampled in coarser beams (i ~ 36° outside 70 AU; Piétu, Guilloteau, & Dutrey Reference Piétu, Guilloteau and Dutrey2005). Hashimoto et al. (Reference Hashimoto2011) also conclude, from polarised-differential-imaging (PDI) in the near-IR, that the inner regions of AB Aur are warped, given the varying inclinations in a double ring structure, dropping from ~ 43° to ~ 27° over 90 to 200 AU in radius. Tang et al. (Reference Tang, Guilloteau, Piétu, Dutrey, Ohashi and Ho2012) relate the warped structure inferred from their molecular line data with a residual infalling envelope.

Yet, another example of such inclination changes with angular scale is seen in MWC 758 (~ 2 M), where near-IR interferometry (with VLTI+AMBER, Isella et al. Reference Isella, Tatulli, Natta and Testi2008) yields an inclination of 30°–40° in the central AU, while Submillimeter Array (SMA) observations in CO(3 − 2), with a 0.7 arcsec beam spanning over 100 AU, yield an inclination of ~ 21° (Isella et al. Reference Isella, Natta, Wilner, Carpenter and Testi2010). Similarly, in HD 135344B (also ~ 2M), the VLTI+MIDI inclination is ~ 60° (Fedele et al. Reference Fedele2008), while CO(3 − 2) data suggests 11° (Dent, Greaves, & Coulson Reference Dent, Greaves and Coulson2005), and IR direct imaging restricts < 20° (HST+NICMOS; Grady et al. Reference Grady2009).

From an observational point of view, there is a degeneracy in the line-of-sight kinematics due to variable inclination, as in a warp, and the non-Keplerian flows expected from infalling gas, as identified in TW Hya and HD 142527 by Rosenfeld et al. (Reference Rosenfeld2012) and Rosenfeld, Chiang, & Andrews (Reference Rosenfeld, Chiang and Andrews2014). In TW Hya, excess CO emission at high velocities, additional to that expected from axially symmetric disk models, could be due to a warp, that may also account for the faint azimuthal modulation in scattered light imaging reported by Roberge, Weinberger, & Malumuth (Reference Roberge, Weinberger and Malumuth2005). In HD 142527, the infalling gas reported by Casassus et al. (Reference Casassus2013b) and Casassus et al. (Reference Casassus, Perez, van der Plas, Dent, Hales and Ménard2013a) could instead be due to a warp.

2.3. The HD 142527 warp

The unambiguous identification of warps requires further evidence, in addition to the possible inclination trends with angular scales (which could be related to a variety of physical structures other than warps) and the hints provided by the molecular line data. Radiative transfers effects due to warps can provide such evidence. For instance, in HD 100546, Quillen (Reference Quillen2006) recognised how a warped structure could approximate the spiral pattern seen by HST (Grady et al. Reference Grady2001), thereby illustrating the potential of radiative transfer effects to account for non-axially symmetric structures. Indeed, Whitney et al. (Reference Whitney, Robitaille, Bjorkman, Dong, Wood and Honor2013) apply three-dimensional radiative transfer to warped disk architectures, as shown in Figure 1, showing that shadows are expected in the outer disks.

Figure 1. Figure adapted from Whitney et al. (Reference Whitney, Robitaille, Bjorkman, Dong, Wood and Honor2013, ©AAS, reproduced by permission), with a three-dimensional radiative transfer prediction in JHK for a warped disk configuration. The field is 4 AU on a side.

HD 142527 is an example of how such radiative transfer effects in scattered light can unambiguously determine the disk orientation and the existence of variable inclinations. In this case, the relative inclination change between the outer and inner disks reaches ~ 70° ± 5°, as shown by Marino, Perez, & Casassus (Reference Marino, Perez and Casassus2015a). Despite such a dramatic inclination change, the intensity dips originally identified by Casassus et al. (Reference Casassus, Perez, Jordán, Ménard, Cuadra, Schreiber, Hales and Ercolano2012) eluded interpretation for 3 yrs, because the direction connecting the dips is offset from the star, in contrast with that naively expected for a simple tilt. Yet, as illustrated in Figure 2, the silhouette of the shadows cast by the inner warp provide unequivocal evidence for this warp, which is further supported by the gas kinematics discussed in Section 3.

Figure 2. Comparison between the observed H-band polarised intensity image of HD 142527 (a- from Avenhaus et al. Reference Avenhaus, Quanz, Schmid, Meyer, Garufi, Wolf and Dominik2014a) and three-dimensional radiative transfer predictions (b- from Marino et al. Reference Marino, Perez and Casassus2015a, this is an updated version of their Figure 2). The kinematics of C18O(2 − 1) emission (Perez et al. Reference Perez2015a) give the orientation of the outer disk; the white contours in a- correspond to systemic velocities, so that the PA of the outer disk lies at ~ 160° East of North, as indicated on b-. The inner disk shadows cast on the outer disk are best reproduced with a PA of 172°, the curvature of their outline (or silhouette) is reminiscent of the observations for the Eastern side. The similarities with the observations are particularly good given the idealisations of the model, which assumes a circular cavity.

2.4. Near-term prospects: how common are warps?

Although there are indications for the frequent occurrence of inclination changes in gas-rich systems, there is so far only one clear case of a warp, i.e. in HD 142527. Next-generation adaptive-optics cameras, such as SPHERE and GPI, should soon provide substantial improvements, particularly through PDI which appears ideally suited to trace illumination effects, and so identify warps. First, results from such new AO imaging are indeed revealing intriguing intensity modulations, as in MWC 758 (Benisty et al. Reference Benisty2015), HD 135344B (Wahhaj et al. Reference Wahhaj2015), and in HD 100453 (Wagner et al. Reference Wagner, Apai, Kasper and Robberto2015). Follow-up molecular-line observations with Atacama Large Millimeter Array (ALMA) may provide the necessary clues from the gas kinematics.

3 GAS IN CAVITIES

Models predict that young giant protoplanets carve a deep gap in the dust component of protoplanetary disks, and a shallower gap in the gas and very small grains (VSGs) components (e.g. Paardekooper & Mellema Reference Paardekooper and Mellema2006; Fouchet, Gonzalez, & Maddison Reference Fouchet, Gonzalez and Maddison2010). The clearing of the protoplanetary gap is thought of as an important mechanism underlying the class of TDs, and residual gas in cavities has thus been intensely searched for.

This section describes observations of intra-cavity gas that select sufficiently large cavities to be resolved with current gas tracers, so with radii greater than a dozen AUs at the very least. Smaller cavities and inner disks have been detected by broad-band long-baseline optical interferometry (e.g. the gaps in HD 100546 or HD 139614 [Tatulli et al. Reference Tatulli2011, Matter et al. Reference Matter2016]). Such observations do not yet sample the intra-cavity gas (except for the central few stellar radii in Brɣ, e.g. Mendigutia et al.). The continuum interferometric observations have allowed to isolate the crystalline spectrum of the inner disks (as in HD 142527 [van Boekel et al. Reference van Boekel2004]), and to study the occurrence bof gaps in HAeBes (Menu et al. Reference Menu, van Boekel, Henning, Leinert, Waelkens and Waters2015)

3.1. Ro-vibrational gas and very small grains

Long-slit spectroscopy has provided indirect evidence for gas inside cavities, under the assumption of azimuthal symmetry and Keplerian rotation (Carr, Mathieu, & Najita Reference Carr, Mathieu and Najita2001; Najita, Carr, & Mathieu Reference Najita, Carr and Mathieu2003; Acke & van den Ancker Reference Acke and van den Ancker2006; van der Plas et al. Reference van der Plas, van den Ancker, Fedele, Acke, Dominik, Waters and Bouwman2008; Salyk et al. Reference Salyk, Blake, Boogert and Brown2009). Spectro-astrometry has also been used to infer a residual gas mass from ro-vibrational CO emission (Pontoppidan et al. Reference Pontoppidan, Blake, van Dishoeck, Smette, Ireland and Brown2008; van der Plas et al. Reference van der Plas, van den Ancker, Acke, Carmona, Dominik, Fedele and Waters2009; Pontoppidan, Blake, & Smette Reference Pontoppidan, Blake and Smette2011). A comprehensive analysis is presented in van der Plas et al. (Reference van der Plas, van den Ancker and Waters2015) and Banzatti & Pontoppidan (Reference Banzatti and Pontoppidan2015). The general picture is that dust cavities do indeed contain some amount of residual gas, as expected in the context of dynamical clearing.

The smallest grains are thought to be well coupled with the gas, so that they can be used as proxies. Thermal IR observations suggest the existence of smaller gaps in VSGs than in silicates, and have also led to the finding of gaps otherwise undetectable through the SED alone (Maaskant et al. Reference Maaskant2013; Honda et al. Reference Honda2015). Detailed studies of the gap structure in ro-vibrational lines along with infrared data have been performed in HD 135344B (Garufi et al. Reference Garufi2013; Carmona et al. Reference Carmona2014).

3.2. Resolved radio observations of intra-cavity gas

While the ro-vibrational detections pointed at residual intra-cavity gas, surprises were nonetheless brought by the first resolved images (Figure 3, Casassus et al. Reference Casassus2013b), made possible thanks to the advent of ALMA. The finest angular resolutions from ALMA Cycle 0 were just sufficient to resolve the largest TD cavity, i.e. that of the HD 142527 disk. A few M jup worth of gas was indeed seen inside this cavity from CO isotopologues (Perez et al. Reference Perez2015a). However, the gas kinematics were puzzling, and seemed consistent with radial infall as illustrated by the fast and centrally peaked HCO+(4 − 3). Slower intra-cavity HCO+ signal seemed to connect with the outer disk along filamentary structures, reminiscent of protoplanetary accretion streams. But the slow intra-cavity HCO+ was faint in these Cycle 0 data, and the cavity radius was sampled only with three beams. New observations in HCO+(4 − 3) are required to ascertain its structure.

Figure 3. Summary of Cycle 0 band 7 observations, from MEM maps, with continuum in red, HCO+(4 − 3) in green, and CO(3 − 2) in blue (adapted from Casassus et al. Reference Casassus2013b). x − and y − show offset from the star, in arcsec. Velocities have been restricted to highlight the fainter structures seen in HCO+, which are otherwise dwarfed by the fast HCO+ central emission.

A more accurate view of the intra-cavity kinematics in HD 142527 is provided by the CO(6 − 5) line, with a higher frequency and a finer Cycle 0 beam. CO(6 − 5) is also free of the interstellar absorption that affects CO(3 − 2) (Casassus et al. Reference Casassus2013c, Reference Casassus, Perez, van der Plas, Dent, Hales and Ménard2013a). Given the warped structure of HD 142527, the line-of-sight CO(6 − 5) can be understood as infall, at the observed stellar accretion rate (Garcia Lopez et al. Reference Garcia Lopez, Natta, Testi and Habart2006), but along and through the warp. As illustrated in Figures 4 and 5, the observations are fairly well accounted for by a model with gas inside the continuous warp. At the radius where the disk plane crosses the sky, the kinematics are Doppler-flipped so that blue turns to red, and convolution with the angular resolution results in the ‘S’-shaped centroids. Interestingly, a fast and narrow warp improves the model, such that the two disk orientations are connected within ~ 3 AU, at a radius of 20 AU, with material flowing orthogonal to the local disk plane at a velocity comparable to Keplerian. These observations are based on non-parametric image synthesis that pushes the limits of the Cycle 0 beam, so that further observations are required to ascertain the details of the gaseous flows inside the warp.

Figure 4. Figure adapted from Casassus et al. (Reference Casassus2015a), showing the observed CO(6 − 5) kinematics in the central regions of HD 142527, and comparing with model predictions for the accretion through the warp represented in Figure 5 (convolved at the resolution of the centroid map). The origin of coordinates is set to the stellar position.

Figure 5. Sketch of the accretion kinematics through the warp in HD 142527 (from Casassus et al. Reference Casassus2015a). The crescent in hues of red and purple represents the distribution of mm-sized grains. Inside the warp, as it connects the two disk inclination material flows along the dashed and curved arrows.

A likely origin for the observed accretion dynamics in HD 142527 probably involves the low-mass companion, at ~ 12 AU and with a mass ratio ≲ 1/10 (Biller et al. Reference Biller2012; Close et al. Reference Close2014; Rodigas et al. Reference Rodigas, Follette, Weinberger and Close2014). The proper-motion of the companion (Lacour et al. Reference Lacour2015) seems to indicate an orbit in clockwise rotation, as for the whole disk. If HD 142527B is contained in the tilted inner disk, or if it is close to it, the properties of the warp caused by the HD142527A+B system resemble disk tearing in strongly warped circumbinary disks (Nixon, King, & Price Reference Nixon, King and Price2013; Nealon, Price, & Nixon Reference Nealon, Price and Nixon2015; Doğan et al. Reference Doğan, Nixon, King and Price2015). In the context of circumstellar disks, Facchini, Lodato, & Price (Reference Facchini, Lodato and Price2013) and Facchini, Ricci, & Lodato (Reference Facchini, Ricci and Lodato2014) show that in the thick-disk regime, where α < H/R, an inner warp greater than ~ 40° will break the disk into distinct planes. Note, however, that the low mass companion at ~ 12 AU is unlikely to have originated the 140 AU cavity, unless it is undergoing oscillations in inclination and eccentricity (i.e. Kozai cycles, see Martin et al. 2014). Further progress in the interpretation of the HD 142527 warp requires better data and tailored hydrodynamical models.

3.3. Residual gas in cavities

A summary of several other detections with ALMA of residual gas in TD cavities is given by van Dishoeck et al. (Reference van Dishoeck, van der Marel, Bruderer and Pinilla2015). Bright CO(6 − 5) emission was found ascribed to the cavities of HD 135344B and SR21 (Pérez et al. Reference Pérez, Isella, Carpenter and Chandler2014), as well as in LkCa 15, RXJ 1615-3255, and SR 24S (van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Pérez and Isella2015a). The size of the ring in [PZ99]J160421.7-213028 is smaller in CO(2 − 1) than in the dust continuum (Zhang et al. Reference Zhang, Isella, Carpenter and Blake2014), much as in SZ 91(Canovas et al. Reference Canovas2015). The depth of the cavity in mm-sized grains follows from the submillimetre continuum. Results so far are limited by dynamic range, with depths in the mm-sized dust deeper than ~ 1/100 relative to the outer disk.

The recent state of the art 345 GHz continuum and CO(3 − 2) isotopologue data analysed by van der Marel et al. (Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016) in terms of a thermochemical model, suggests that the gas cavities are up to three times smaller than the dust rings in DoAr 44, SR 21, HD 135344B, and IRS 48 (see also Bruderer et al. Reference Bruderer, van der Marel and van Dishoeck2014). The drop in gas surface density can be up to δ ~ 10−2 (the exact values vary in each object and can be calculated from the parameters given in Table 3 of van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016).

The main result that transpires from the above studies, including HD 142527 (Perez et al. Reference Perez2015a), is that the depth of the cavity is shallower in the gas than in the dust, by at least a factor of 10. However, while van der Marel et al. (Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016) investigate gradual cavity edges, angular resolution limits prevent firm conclusions on the distribution of the gas. Most studies assume axially symmetric and sharp cavity edges, parametrised by a step-function drop, with a depth δ ~ 1/10 − 1/100 relative to the outer disk in the gaseous component. Due to the coarse linear resolutions available, the accretion kinematics remain largely unsampled. For instance, in HD 142527 sufficient linear resolution is crucial to trace the non-Keplerian accretion flows, which are otherwise much less conspicuous when sampled with a beamwidth comparable to the cavity radius (Perez et al. Reference Perez2015a).

3.4. Near-term prospects: structure and origin of the cavities

The finest angular resolutions available from the latest ALMA results (van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016) do not yet provide enough linear resolution to sample the cavity edges. However, at the time of writing several ALMA projects are currently underway in gas-rich cavities, and the first results from the next-generation AO cameras are soon due. With much finer angular resolutions, it will be possible to understand better the distribution of the gas in TDs and its kinematics, and so reconcile mass transport across TD cavities with the stellar accretion rates (which has so far only been possible in HD 142527).

A key question is on the shape of the cavity edge, which is currently modelled as a step function. Yet, the edge of the cavity gives important clue on the clearing mechanism: for instance, if planet formation, the sharpness of the edge is a function of the mass of the outermost body (e.g. Crida, Morbidelli, & Masset Reference Crida, Morbidelli and Masset2006; Mulders et al. Reference Mulders, Paardekooper, Dominik, van Boekel and Ratzka2013).

Intriguingly very few companions have been found in TDs (and only one candidate near the edge of a cavity, i.e. in the HD 169142 disk, see Section 6). A spectacular example is HD 142527, with a record-sized cavity of 140 AU and no intra-cavity protoplanet detected as yet. The binary separation in HD 142527 is ~ 10 AU (see Section 3.2) and could only account for a ~ 30 AU gas-free cavity (Artymowicz & Lubow Reference Artymowicz and Lubow1994). An interesting comparison object is the GG Tau circumbinary ring, which by contrast to HD 142527 may be explained in terms of dynamical interactions with the binary provided the edge of the cavity is ‘soft’ in the gas (Andrews et al. Reference Andrews2014; Dutrey et al. Reference Dutrey2014).

Despite the lack of detections in TD cavities, the data in HD 142527 suggest that low-mass companions can drive dramatic warps. Thus, a possible solution to the puzzle of large cavities could perhaps be found in viewing them as circumbinary disks with low mass companions undergoing Kozai oscillations, so that when the system is coplanar the binary is very eccentric (as suggested in Casassus et al. Reference Casassus2015a, pending tests with hydrodynamic simulations). Thus, in HD 142527, we could be witnessing a high-inclination phase at relatively close binary separation. Definitive conclusions on the structure of this warp require finer angular resolutions and deeper sensitivities, along with new observations in the HCO+ lines. Here, we are presented with another puzzle to reconcile models of disk tearing with the indications that even though the warp appears to be abrupt, gas continuously flows through it and connects the two disk inclinations.

Is there a binary lurking in all of the large TD cavities, as in HD 142527? Low mass companions into the brown-dwarf mass regime are difficult to detect at very close separations. The new AO cameras along with ALMA gas kinematics should soon cast light on these questions, starting with the recent detection of fairly high-mass bodies in LkCa 15 (Sallum et al. Reference Sallum2015, see Section 6).

4 LOPSIDED DISKS

Early resolved observations of TDs already suggested that the outer rings deviate from axial symmetry (in general from SMA data, Ohashi Reference Ohashi2008; Brown et al. Reference Brown, Blake, Qi, Dullemond, Wilner and Williams2009; Andrews et al. Reference Andrews, Wilner, Espaillat, Hughes, Dullemond, McClure, Qi and Brown2011). ALMA has brought confirmation and revealed the dramatic contrastFootnote 4 of the most extreme asymmetries seen in the dust continuum. To what extent do these asymmetries trace matched asymmetries in the gas? What is the efficiency of dust segregation from the gas?

4.1. The most dramatic asymmetries

4.1.1. HD 142527

The submillimetre continuum from HD 142527 is approximately shaped into a crescent (Ohashi Reference Ohashi2008; Casassus et al. Reference Casassus2013b; Fukagawa et al. Reference Fukagawa2013) and modulated by a temperature decrement under the shadow of the inner warp (Section 2.3). It is still not entirely clear to what extend is the submillimetre continuum reflecting a similarly pronounced asymmetry in the gas. Existing observations rely on optically thick tracers of the total gas mass (including small dust grains), so cannot conclude on the underlying gas distribution. The dense-gas tracers discussed by van der Plas et al. (Reference van der Plas, Casassus, Perez, Thi, Pinte and Christiaens2014, HCN(4 − 3) and CS(7 − 6)) are affected by chemistry and seem to anti-correlate with the continuum.

Recently, the Cycle 0 data in the optically thin CO isotopologues were further analysed by Muto et al. (Reference Muto2015), who estimate that the contrast in the gas-phase CO column is about a factor of 3. While this is all the existing information available on the gas contrast, at the time of writing, its extrapolation to the total gas mass is hampered by uncertainties from grain surface chemistry: The azimuthal structure in these transitions is not a smooth crescent; their emission peaks differ from the continuum. Perhaps, the gas-phase CO abundance is significantly modulated in azimuth by chemistry, or a fraction of CO could be depleted on dust grains, especially at the location of the continuum peak (at about 11 h), where continuum grey-body temperatures are ~ 22 K (Figure 6, Casassus et al. Reference Casassus2015d), so below freeze-out (Jørgensen et al. Reference Jørgensen, Visser, Williams and Bergin2015).

Figure 6. Figure adapted from Casassus et al. (Reference Casassus2015d). Upper row: multi-frequency data of HD 142527 brought to a common uv-coverage. The ALMA data have been filtered for the ATCA response with Monte Carlo simulations of ATCA observations on deconvolved models of the ALMA data. (a): restored ATCA image at 34 GHz. (b): average of Monte Carlo (MC) simulations of ATCA observations on the ALMA band 7 data. (c): MC simulations of ATCA observations on the ALMA band 9 data. Lower row: Grey-body diagnostics inferred from the multi-frequency data. (a): optical depth map at the reference frequency of 345 GHz. (b): line of sight emissivity index map β, with ATCA specific intensity contours in red. (c): root-mean-square uncertainties on the emissivity index map. (d): line of sight temperature, Ts .

The hydrodynamic simulations tend to predict contrast ratios of order ~ 3. Large-scale crescents in the gas surface density arise naturally in models of cavity clearing by giant planets (e.g. Zhu & Stone Reference Zhu and Stone2014). Such crescents are reproduced in hydrodynamical simulations by large-scale anti-cyclonic vortices, which result from Rossby wave instabilities triggered at sharp radial gradients in physical conditions. For instance, they have also been modelled in the context of viscosity gradients (Regály et al. Reference Regály, Juhász, Sándor and Dullemond2012). An apparently different model for such crescents, leading to more pronounced asymmetries in the gas, was proposed by Mittal & Chiang (Reference Mittal and Chiang2015) based on a global disk mode in response to a stellar offset. However, as argued by Zhu & Baruteau (Reference Zhu and Baruteau2015), if such offsets result from the shift of the system centre of mass in sufficiently massive disk, the consistent incorporation of disk self-gravity dampens the contrast of an otherwise standard large-scale anticyclonic vortex. Thus in general, the gas surface density contrasts are predicted to reach moderate values, typically ~ 3 and perhaps up to 10.

The azimuthal intensity contrast in HD 142527 reaches about 30, yet there is no hint so far of a similar contrast ratio either in the small grains (≪ 0.1 mm), nor in molecular lines, which led Casassus et al. (Reference Casassus2013b) to suggest that the mm-sized dust grains that originate the ALMA continuum were segregated from the gas. In other words, the dust-to-gas mass ratio could vary with azimuth. A promising mechanism to explain such segregation is the pile-up of larger grains in local pressure maxima (e.g. Weidenschilling Reference Weidenschilling1977; Barge & Sommeria Reference Barge and Sommeria1995; Birnstiel, Dullemond, & Pinilla Reference Birnstiel, Dullemond and Pinilla2013; Lyra & Lin Reference Lyra and Lin2013), as proposed to account for the even more dramatic submillimetre continuum contrast seen in IRS 48 (see below, van der Marel et al. Reference van der Marel2013).

An important prediction of dust trapping is that progressively larger grainsFootnote 5 should be more sharply confined. Multi-frequency radio observations may test this prediction, since grains of progressively larger size dominate the continuum emission at correspondingly longer wavelengths. Indeed, 34 GHz ATCA observations of HD 142527 revealed a compact clump of cm-wavelength-emitting grains buried into the ALMA crescent (Figure 6, Casassus et al. Reference Casassus2015d). Populations of larger grains correspond to shallower optical depth spectral index β (e.g. Testi et al. Reference Testi2014), in a parametrisation of the optical depth spectrum such that τ = τ × (ν/ν)β s (β is also referred to as emissivity index). Thus, three frequency points can be used to solve for the radiation temperature, optical depth, and β index at each line of sight. Frequencies above ~ 345 GHz turned out to be optically thick (Figure 6, Casassus et al. Reference Casassus2015d), so that the spectral index trends at ALMA frequencies are due to optical depth effects rather than the underlying dust population. The ATCA clump translates into a local minimum in the β index, indicative of larger grains.

Interestingly, the optical depth at 345 GHz (Figure 6 a, bottom), shows an extension towards 2 h. Yet, the minimum in opacity spectral index β lies at 11 h, which should be the core of the pressure maximum in the dust trap scenario. Since rotation in HD 142527 goes clockwise, a possible interpretation lies in the phenomenon predicted by Baruteau & Zhu (Reference Baruteau and Zhu2015), that larger grains should concentrate significantly ahead of the dust trap. Thus at 2 h, the line of sight would intercept a disk-averaged grain population, modulated by the gaseous crescent, but the addition of larger grains (perhaps up to 1 cm sizes in this disk), might raise the total optical depth even away from the pressure maximum. However, such a shift of the larger grains is possible only with Stokes numbers S ≳ 1 (Baruteau & Zhu Reference Baruteau and Zhu2015), which for 1~cm grains would require a disk mass ~ 100 times lower than inferred from the SED modelling (Casassus et al. Reference Casassus2015d), and/or a correspondingly low gas-to-dust mass ratio.

Thus, the bulk of the data so far indicates that the spatial segregation of dust sizes in azimuth is indeed at work in the lopsided outer ring of HD 142527. However, while models have been proposed that approximate the observed segregation (through parametric models, Casassus et al. Reference Casassus2015d), they are fine-tuned in key parameters such as the α turbulence prescription and the threshold grain size for trapping.

4.1.2. IRS 48

The record-holder lopsided TD is IRS 48, where van der Marel et al. (Reference van der Marel2013) found that the contrast in ALMA band 9 is greater than 100, so three times as pronounced as in HD 142527. They reproduce such dramatic contrast with their dust trapping prescriptions.

Observations of IRS 48 are hampered by intervening cloud emission from ρ Oph, with A V ~ 10. Yet, by using the rarer CO isotopologues at higher velocities, Bruderer et al. (Reference Bruderer, van der Marel and van Dishoeck2014) estimated that the total gas mass of the disk is only ~ 10−4M.

The dramatic contrast observed in IRS 48 would seem very unlikely to correspond to an equally lopsided gas distribution, and is thus probably due to segregation of grain sizes. Indeed, van der Marel et al. (Reference van der Marel, Pinilla, Tobin, van Kempen, Ricci and Birnstiel2015b) report on the tentative detection of the corresponding spectral trends in a comparison between VLA and ALMA observations. The confirmation of lopsidedness at optically thin VLA frequencies (see Figure 7) dissipates worries that the asymmetry observed at high frequencies, which are likely optically thick, could be related to optical depth effects in this highly inclined disk.

Figure 7. Figure adapted from van der Marel et al. (Reference van der Marel, Pinilla, Tobin, van Kempen, Ricci and Birnstiel2015b, their Figure 1, ©AAS, reproduced with permission). VLA observations of IRS 48 at optically thin frequencies confirm the extreme lopsidedness of this TD. VLA observations at 34 GHz are shown in red contours—the tail towards the NE is likely due to stellar emission. ALMA band 9 observations at 680 GHz are shown in black contours, after filtering for the VLA response. The field is 2 arcsec on a side. While at 34 GHz the crescent appears to be somewhat more compact, the role of optical depth effects in widening the 680 GHz signal remain to be quantified.

4.1.3. MWC 758

Can all continuum asymmetries in TDs be interpreted as dust traps? MWC 758 presents an interesting caveat. The VLA 34 GHz image presented by Marino et al. (Reference Marino, Casassus, Perez, Lyra, Roman, Avenhaus, Wright and Maddison2015b, reproduced in Figure 8), when compared with the more extended signal seen by ALMA, seems fairly consistent with the dust trap scenario, at least as well as in HD 142527 or IRS 48 given the available constraints. But in fact two pressure maxima (or anticyclonic vortices) would seem to be required to account for the double-peaked ALMA signal in MWC 758. Besides estimates of the submillimetre optical depth are missing—which requires finer angular resolution submillimetre data, as well as an additional frequency point intermediate with the VLA.

Figure 8. Figure adapted from Marino et al. (Reference Marino, Casassus, Perez, Lyra, Roman, Avenhaus, Wright and Maddison2015b, their Figure 3). VLA observations of MWC 758 compared to a deconvolved model of the ALMA band 7 visibilities. The VLA signal is shown in red contours—the red ellipse corresponds to the synthesised beam. The blue scale corresponds to an ‘MEM’ non-parametric model of the 337 GHz ALMA data—the blue ellipse corresponds to an elliptical Gaussian fit to the point-spread-function of the ‘MEM’ algorithm.

The compact VLA signal in MWC 758 may also be looked at from an entirely different perspective. Another interpretation could follow from the recent proposal by Dong et al. (Reference Dong, Zhu, Rafikov and Stone2015a) that the observed spiral pattern in MWC 758 are launched by planet could be launched by a planet exterior to the arms, at a radius of ~ 0.6 arcsec. Marino et al. (Reference Marino, Casassus, Perez, Lyra, Roman, Avenhaus, Wright and Maddison2015b) caution that the observed signal is quite close to the location of such a body, as illustrated in Figure 9. Thus, the compact signal could be circumplanetary dust that is heated by planet formation feedback, (as in the positive feedback leading to enhanced viscous heating proposed by Montesinos et al. Reference Montesinos, Cuadra, Perez, Baruteau and Casassus2015).

Figure 9. Left: Spiral model involving a planetary mass companion, from Dong et al. (Reference Dong, Zhu, Rafikov and Stone2015a, part of their Fig. 4, ©AAS, reproduced with permission). The companion is highlighted as a green dot—and the field is rotated so that the companion approximately matches the location of the VLA clump seen in MWC 758. Right (adapted from Marino et al. Reference Marino, Casassus, Perez, Lyra, Roman, Avenhaus, Wright and Maddison2015b): VLA data, in red contours, and ALMA 345GHz, in blue contours, overlaid on the SPHERE Y-band polarised intensity image (Benisty et al. Reference Benisty2015), in grey scale.

4.2. Mild asymmetries and radial dust traps

The above examples for extreme submillimetre continuum asymmetries are seen in few HAeBe stars (stellar mass 2 M), and are not typical of the average TD. Most TDs show milder azimuthal asymmetries when resolved in the submillimetre continuum. Examples of mild contrast ratios, ≲ 4 or as could be reached by the gas, are HD 135344B, SR 21 (Pérez et al. Reference Pérez, Isella, Carpenter and Chandler2014; Pinilla et al. Reference Pinilla2015), and LkHα 330 (Isella et al. Reference Isella, Pérez, Carpenter, Ricci, Andrews and Rosenfeld2013). Given their coarseness, these observations could accommodate sharper asymmetries.

An interesting avenue to explain such mild asymmetries could be spiral arm crowding. Indeed the asymmetry reported in HD 135344B (Pérez et al. Reference Pérez, Isella, Carpenter and Chandler2014) with a relatively coarse beam in ALMA Cycle 0, has been resolved into a convergence of spiral arms (van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016, their Figure 1). These spirals were hinted at in axisymmetric model residuals (Pérez et al. Reference Pérez, Isella, Carpenter and Chandler2014; Pinilla et al. Reference Pinilla2015). Local shock heating from the spiral waves (Rafikov Reference Rafikov2016) could perhaps further pronounce asymmetries related to spirals, and lead to mild asymmetries when observed in a coarse beam.

There are also examples of axially symmetric TDs, such as the TTauri J160421, SZ 91, DoAr 44 (Zhang et al. Reference Zhang, Isella, Carpenter and Blake2014; van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016), and SZ 91 (Canovas et al. Reference Canovas, Caceres, Schreiber, Hardy, Cieza, Ménard and Hales2016). While azimuthal dust segregation or trapping does not seem to be required in such systems with no or mild asymmetries, on the other hand, the radial segregation of dust grain sizes seems to be systematically present in TDs. Radial trapping is often invoked as a means to halt the catastrophic inward migration of grains due to aerodynamic drag (Weidenschilling Reference Weidenschilling1977). The smaller and shallower cavities in the gas phase compared to the dust were introduced in Section 3 as a feature of dynamical clearing by giant planet formation. This is the same phenomenon as radial trapping, since it is the radial bump in gas pressure generated by planet formation that traps the larger dust grains in the outer disk (e.g. Pinilla, Benisty, & Birnstiel Reference Pinilla, Benisty and Birnstiel2012; Pinilla et al. Reference Pinilla2015).

4.3. Near-term prospects: the gas background and vortex velocity fields

The dust trap phenomenon is clearly at work in the two most extreme asymmetries, and could be an important mechanism in the evolution of the dust grain population. However, accurate knowledge of the gas background and physical conditions is required to estimate the efficiency of trapping by aerodynamic coupling. As yet, there has been no confirmation of the velocity fields expected in large-scale vortices, so an important unknown on the origin of the lopsided gas distributions remains untackled. Another open question is, for instance, the impact of enhanced cooling in the dust trap on the maximum possible gas density contrast.

Taking the dust trapping scenario one step further poses a question on the potential of such dust traps to form gaseous giants. What would be the impact on the disk of massive bodies forming inside a dust trap at large stellocentric distances? Are there signs of such bodies?

Both the observational and theoretical situations are quickly evolving. ALMA should soon provide additional multi-frequency continuum data with which to better quantify the degree of trapping and the corresponding grain distributions. Perhaps, resolved molecular line maps of dust traps could be used to detect the expected vortex velocity fields. Without an observational confirmation, however, the question on the origin of the gaseous lopsided structures is still open.

5 SPIRALS

Observations in optical/IR scattered light of spirals in protoplanetary disks have been discussed on several occasions, both in discovery articles (see below) and in modelling efforts (Juhász et al. Reference Juhász, Benisty, Pohl, Dullemond, Dominik and Paardekooper2015; Pohl et al. Reference Pohl, Pinilla, Benisty, Ataiee, Juhász, Dullemond, Van Boekel and Henning2015; Dong et al. Reference Dong, Hall, Rice and Chiang2015b). The single-star systems where spirals have been reported seem to be TDs, i.e. they host either central cavities or gaps. It is possible that these systems are in fact binaries with a low mass ratio: their spiral patterns bear similarities with the arc-like features seen in some binary disks, such as in AS 205 (Salyk et al. Reference Salyk, Pontoppidan, Corder, Muñoz, Zhang and Blake2014) and SR 24 (Mayama et al. Reference Mayama2010). The following is a list of spiral patterns in disks around primary stars that dominate by mass (relative to possible low mass bodies embedded in the disks), some of which are shown in Figure 10:

Figure 10. Gallery of optical/IR spirals. From left to right, we show scattered-light images of AB Aur (Fukagawa et al. Reference Fukagawa2004, part of their Fig. 3, ©AAS, reproduced with permission), HD 141569A (Clampin et al. Reference Clampin2003, part of their Fig. 8, ©AAS, reproduced with permission), HD 135344B (Garufi et al. Reference Garufi2013, part of their Fig. 1, ©ESO, reproduced with permission), and HD 100453 (Wagner et al. Reference Wagner, Apai, Kasper and Robberto2015, part of their Fig. 2, ©AAS, reproduced with permission). All figures have been reproduced by permission of the AAS or A&A. In HD 100453, we have highlighted two intensity dips where the two-armed spiral pattern seem to stem from. These dips are very reminiscent of the HD 142527 shadows, which also seem to be at the root of spirals.

There are, however, only two examples of molecular-line counterparts to the near-IR spirals in protoplanetary disks. The AB Aur IR spirals were also detected in the CO isotopologues (Corder, Eisner, & Sargent Reference Corder, Eisner and Sargent2005; Lin et al. Reference Lin, Ohashi, Lim, Fukagawa and Tamura2006). Intriguingly, the spirals in AB Aur appear to counter-rotate with the disk, which led Tang et al. (Reference Tang, Guilloteau, Piétu, Dutrey, Ohashi and Ho2012) to propose that they stem from an infalling envelope (but the apparent counter-rotation could also result from a Doppler flip due to a warp, see Section 2.2). As summarised in Figure 11, Christiaens et al. (Reference Christiaens, Casassus, Perez, van der Plas and Ménard2014) reported very large scale 12CO spirals in HD 142527, one of which appears to have a clear counterpart in the most conspicuous IR spiral found by Fukagawa et al. (Reference Fukagawa, Tamura, Itoh, Kudo, Imaeda, Oasa, Hayashi and Hayashi2006).

Figure 11. Radio spirals from HD 142527 (Figure adapted from Christiaens et al. Reference Christiaens, Casassus, Perez, van der Plas and Ménard2014). The CO(2 − 1) peak signal is shown on the left, while CO(3 − 2) is shown on the right (the gray ellipse correspond to the clean beams). x − and y − axis indicate angular offset from the star, along RA and Dec., and in arcsec. Since the distance to HD 142527 is about 140 pc, the spirals are seen to extend out to 700 pc in radius. The lozenge symbols indicate the most conspicuous IR spiral, from Fukagawa et al. (Reference Fukagawa, Tamura, Itoh, Kudo, Imaeda, Oasa, Hayashi and Hayashi2006). The inverted-V decrement in peak CO intensity, seen in both transitions, is due to interstellar absorption at velocities that correspond to these locations in the Keplerian outer disk.

The radio detection of spirals in molecular-lines is very important to estimate physical conditions, and search for kinematic trends, such as deviations from Keplerian rotation (as in AB Aur), that may help understand the origin of the spirals. Velocity dispersion maps may perhaps also be used to detect the compression wave. For now, the molecular-line spirals in HD 142527 allow estimates of temperature and column densities. Christiaens et al. (Reference Christiaens, Casassus, Perez, van der Plas and Ménard2014) find that the spirals are surprisingly cold, with 25 K near the continuum ring, and down to 10–15 K further out. The Toomre Q parameter, Q = c sΩ/(πGΣ), for each of the two spirals S 1 and S 2, are 100 < Q S1 < 50000, and 50 < Q S2 < 35000, where the lower bound stems from the continuum non-detection (assuming a gas-to-dust ratio of 100), and the upper bound is set by the assumption of optically thin 12CO. Thus, the spirals seen in the outer disk of HD 142527 appear to be gravitationally stable.

5.1. Near-term prospects: radio counterparts

The near future should see a multiplication of radio counterparts to the IR spirals (e.g. the continuum counterparts in HD 135344B; van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Andrews, Pontoppidan, Herczeg, van Kempen and Miotello2016). Radio observations, especially in molecular lines, are required to estimate physical conditions such as density and temperature, and so find clues as to the origin of spirals. In addition, radio observations should help in disentangling radiative transfer effects, such as proposed by Quillen (Reference Quillen2006).

There are several on-going efforts on the theoretical side, with the coupling of three-dimensional hydrodynamics and three-dimensional radiative transfer applied to planet–disk interaction (e.g. Ober et al. Reference Ober, Wolf, Uribe and Klahr2015; Perez et al. Reference Perez, Dunhill, Casassus, Roman, Szulágyi, Flores, Marino and Montesinos2015b; Dong et al. Reference Dong, Zhu, Rafikov and Stone2015a). However, in addition to the gravitational interaction with massive bodies, either planetary-mass or stellar, as in multiple stellar systems, perhaps other physical mechanisms may also launch spirals in circumstellar disks. One possibility is gravitational instability of the disk, whose observability in scattered light has recently been predicted from three-dimensional radiative transfer and hydrodynamics (Dong et al. Reference Dong, Hall, Rice and Chiang2015b). Another example, motivated by the obvious link of the HD 142527 spirals with the shadows cast on the outer disk by the inner warp (which may also be observed in HD 100453), is the proposal by Montesinos et al. (Reference Montesinos, Perez, Casassus, Marino, Cuadra and Christiaens2016) that spirals could result from the forcing of the temperature field in the outer disk, as it is modulated by variable illumination due to non-axially symmetric shadowing from the inner regions. This idea is illustrated in Figure 12.

Figure 12. Spirals launched from the temperature forcing of the outer disk by shadows projected from a tilted inner disk—as in HD 142527 (and perhaps also in HD 100453). The image shows a snapshot of the density field at 250 orbits, under the effect of shadows aligned in the East–West direction (from Montesinos et al. Reference Montesinos, Perez, Casassus, Marino, Cuadra and Christiaens2016).

6 THE ROLE OF EMBEDDED PROTOPLANETS

Key to understanding the astrophysics of circumstellar disks, and eventually giant planet formation, is definitive information on the protoplanetary systems that these disks potentially host. Three promising protoplanet detections have recently been reported from their thermal signal at ~ 3 μm, in the gapped disks HD 100546 (Quanz et al. Reference Quanz, Amara, Meyer, Kenworthy, Kasper and Girard2013, Reference Quanz, Amara, Meyer, Girard and Kasper2015), HD 169142 (Reggiani et al. Reference Reggiani2014) and in LkCa 15 (Sallum et al. Reference Sallum2015). In the case of LkCa 15, the interpretation in terms of an accreting body is strengthened by the detection of coincident and unresolved Hα emission.

The interpretation of unresolved continuum signals at low contrasts is difficult, as the protoplanet accretion luminosity leads to a positive feedback in stellocentric accretion (Montesinos et al. Reference Montesinos, Cuadra, Perez, Baruteau and Casassus2015), resulting in an extended thermal signal well in excess of genuine photospheric emission, and perhaps also a local enhancement in scale height. An example of such extended near-IR emission, both in thermal emission (e.g. in L band) and in scattered-light (e.g. in K band), could be the spatially extended signal from HD 100546b (Garufi et al. Reference Garufi2016).

In general the morphological details of the extended signal from such candidate circumplanetary regions are very difficult to ascertain because of observational biases, either in the high-contrast techniques or in near-IR interferometry, which are being tackled by technological breakthroughs (i.e. with extreme AO cameras such as GPI and SPHERE, and the next-generation interferometry instruments such as VLTI-MATISSE). Examples of such biases are the radiative transfer effects at finite inclination that lead to asymmetries in otherwise axially symmetric disks, and to similar closure phase signals as binary models (as in FL Cha and T Cha, Cieza et al. Reference Cieza2013; Olofsson et al. Reference Olofsson2013).

A different avenue to detect embedded protoplanets, and estimate their mass, is to identify the corresponding kinematical signature of the circumplanetary disk (CPD). Perez et al. (Reference Perez, Dunhill, Casassus, Roman, Szulágyi, Flores, Marino and Montesinos2015b) study the observability of such CPDs with ALMA using molecular lines, by coupling 3D hydrodynamical simulations and radiative transfer, and after filtering for the instrumental responseFootnote 6 . They predict that the CPD from an accreting gaseous giant should stand out in velocity dispersion maps. In addition the channel maps with disk emission at the position of the protoplanet should show a kink (i.e. a tight local bend), whose amplitude depends on the planet mass. Starting from ALMA Cycle 3, such long baseline observations with ALMA are routinely possible, so that prospects are good for the detection of the kinematical signatures of CPDs.

7 CONCLUSION

While accurate knowledge on the existence of embedded protoplanetary systems is still far from systematically available, resolved observations of TDs reveal surprising structures. The observed phenomena do not readily fit in the picture expected for TDs as sculptures of planet formation. Instead, the basic idea of flat protoplanetary disk system is being challenged by the observations of warps, whose frequency remains to be assessed. Residual gas inside cavities is clearly observed, as expected for dynamical clearing, but the concomitant protoplanetary streamers are still elusive. The observed intra-cavity stellocentric accretion, in the single case where it has been detected, occurs at much higher velocities than expected for planetary accretion streams.

The outer regions of TDs often seem to be non-axially symmetric—quite dramatically so in a couple of instances. This observation could bring confirmation for the scenario of ‘dust trapping’ as a way to circumvent the catastrophic infall of the larger dust grains due to aerodynamical drag. While dust size segregation has clearly been demonstrated to exist, the origin of the required gaseous asymmetry in a large anticyclonic vortex remains to be confirmed observationally. There does not seem to be a clear observational connection between the outer disk asymmetry and on-going planet formation, as illustred by the case of LkCa 15, with a detection of protoplanetary accretion (Sallum et al. Reference Sallum2015), and a seemingly perfectly symmetric outer disk (van der Marel et al. Reference van der Marel, van Dishoeck, Bruderer, Pérez and Isella2015a).

The observations of TDs have also revealed spectacular spiral arms, the origin of which still remains the subject of intense debate. They could be launched by planets, or result from gravitational instability, or be induced from variable illumination (Montesinos et al. Reference Montesinos, Perez, Casassus, Marino, Cuadra and Christiaens2016). Once launched, the frequent observation of spirals in systems with cavities and non-axially symmetric outer disks suggests that spirals may play a crucial role in disk evolution, as argued by Rafikov (Reference Rafikov2016). New data will soon provide estimates of physical conditions with which to test these ideas.

Finally, it appears that physical processes other than planet formation are governing the TD phase of Class II YSOs, since most large cavities seem empty of massive bodies. Perhaps, well identifiable planet formation structures have to be searched for at earlier evolutionary stages, i.e. in Class I disks such as HL Tau (ALMA Partnership et al. 2015; Testi et al. Reference Testi2015; Dipierro et al. Reference Dipierro, Price, Laibe, Hirsh, Cerioli and Lodato2015; Gonzalez et al. Reference Gonzalez, Laibe, Maddison, Pinte and Ménard2015; Pinte et al. Reference Pinte, Dent, Ménard, Hales, Hill, Cortes and de Gregorio-Monsalvo2016).

ACKNOWLEDGEMENTS

We thank the following disk workers for very useful input: Henning Avenhaus, Amelia Bayo, Mark Booth, Valentin Christiaens, Lucas Cieza, Sebastián Marino, Matías Montesinos, Johan Olofsson, Sebastián Pérez, Dave Principe, Nienke van der Marel, Gerrit van der Plas and Matthias Schreiber. Financial support was provided by Millennium Nucleus RC130007 (Chilean Ministry of Economy), and additionally by FONDECYT grant 1130949.

Footnotes

1 I.e. ≲ 1 M stars undergoing accretion.

2 The intersection of the plane of the disk with that of the sky.

3 Assuming Keplerian rotation.

4 The ratio of the intensity extrema along a ring.

5 Up to Stokes numbers of ~ 1.

6 Including real phase noise taken from the HL Tau dataset presented by ALMA Partnership et al. (2015).

References

REFERENCES

Acke, B., & van den Ancker, M. E. 2006, A&A, 449, 267 10.1051/0004-6361:20054330 2006A&A...449..267A Google Scholar
Alencar, S. H. P., et al. 2010, A&A, 519, A88 10.1051/0004-6361/201014184 2010A&A...519A..88A Google Scholar
ALMA Partnership, et al. 2015, ApJ, 808, L3 10.1088/2041-8205/808/1/L3 2015ApJ...808L...3A Google Scholar
Andrews, S. M., Wilner, D. J., Espaillat, C., Hughes, A. M., Dullemond, C. P., McClure, M. K., Qi, C., & Brown, J. M. 2011, ApJ, 732, 42 10.1088/0004-637X/732/1/42 2011ApJ...732...42A Google Scholar
Andrews, S. M., et al. 2014, ApJ, 787, 148 10.1088/0004-637X/787/2/148 2014ApJ...787..148A Google Scholar
Artymowicz, P., & Lubow, S. H. 1994, ApJ, 421, 651 10.1086/173679 1994ApJ...421..651A Google Scholar
Avenhaus, H., Quanz, S. P., Meyer, M. R., Brittain, S. D., Carr J. S., & Najita, J. R. 2014b, ApJ, 790, 56 10.1088/0004-637X/790/1/56 2014ApJ...790...56A Google Scholar
Avenhaus, H., Quanz, S. P., Schmid, H. M., Meyer, M. R., Garufi, A., Wolf, S., & Dominik, C. 2014a, ApJ, 781, 87 10.1088/0004-637X/781/2/87 2014ApJ...781...87A Google Scholar
Banzatti, A., & Pontoppidan, K. M. 2015, ApJ, 809, 167 10.1088/0004-637X/809/2/167 2015ApJ...809..167B CrossRefGoogle Scholar
Barge, P., & Sommeria, J. 1995, A&A, 295, L1 1995A&A...295L...1B Google Scholar
Baruteau, C., & Zhu, Z. 2015, arXiv:1511.034981511.03498 2015arXiv151103498B Google Scholar
Benisty, M., et al. 2015, A&A, 578, L6 10.1051/0004-6361/201526011 2015A&A...578L...6B Google Scholar
Biller, B., et al. 2012, ApJ, 753, L38 10.1088/2041-8205/753/2/L38 2012ApJ...753L..38B CrossRefGoogle Scholar
Biller, B. A., et al. 2015, MNRAS, 450, 4446 10.1093/mnras/stv870 2015MNRAS.450.4446B Google Scholar
Birnstiel, T., Dullemond, C. P., & Pinilla, P. 2013, A&A, 550, L8 10.1051/0004-6361/201220847 2013A&A...550L...8B Google Scholar
Boccaletti, A., Pantin, E., Lagrange, A.-M., Augereau, J.-C., Meheut H., & Quanz, S. P. 2013, A&A, 560, A20 10.1051/0004-6361/201322365 2013A&A...560A..20B Google Scholar
Boccaletti, A., et al. 2015, Nature, 526, 230 10.1038/nature15705 2015Natur.526..230B Google Scholar
Bouvier, J., Grankin, K., Ellerbroek, L. E., Bouy, H., & Barrado, D. 2013, A&A, 557, A77 10.1051/0004-6361/201321389 2013A&A...557A..77B Google Scholar
Brown, J. M., Blake, G. A., Qi, C., Dullemond, C. P., Wilner, D. J., & Williams, J. P. 2009, ApJ, 704, 496 10.1088/0004-637X/704/1/496 2009ApJ...704..496B CrossRefGoogle Scholar
Bruderer, S., van der Marel, N., van Dishoeck, E. F., & van Kempen T. A. 2014, A&A, 562, A26 10.1051/0004-6361/201322857 2014A&A...562A..26B Google Scholar
Canovas, H., Ménard, F., Hales, A., Jordán, A., Schreiber M. R., Casassus, S., Gledhill, T. M., & Pinte, C. 2013, A&A, 556, A123 10.1051/0004-6361/201321924 2013A&A...556A.123C Google Scholar
Canovas, H., Caceres, C., Schreiber, M. R., Hardy, A., Cieza, L., Ménard, F., & Hales, A. 2016, MNRAS, submittedhttp://arxiv.org/abs/1601.06801 Google Scholar
Canovas, H., et al. 2015, ApJ, 805, 21 10.1088/0004-637X/805/1/21 2015ApJ...805...21C Google Scholar
Carmona, A., et al. 2014, A&A, 567, A51 10.1051/0004-6361/201322534 2014A&A...567A..51C Google Scholar
Carr, J. S., Mathieu, R. D., & Najita, J. R. 2001, ApJ, 551, 454 10.1086/320071 2001ApJ...551..454C Google Scholar
Casassus, S., Perez, M.S., Jordán, A., Ménard, F., Cuadra, J., Schreiber, M. R., Hales, A. S., & Ercolano, B. 2012, ApJ, 754, L31 10.1088/2041-8205/754/2/L31 2012ApJ...754L..31C Google Scholar
Casassus, S., Perez, S. M., van der Plas, G., Dent, W. R. F., Hales, A., & Ménard, F. 2013a, “Transformational Science with ALMA: From Dust to Rocks to Planets”, NRAO workshop (Waikoloa, Hawaii, April 8–12)pdf/S1-P2_hd142.pdf Google Scholar
Casassus, S., et al. 2015a, ApJ, 811, 92 10.1088/0004-637X/811/2/92 2015ApJ...811...92C Google Scholar
Casassus, S., et al. 2013b, Nature, 493, 191 10.1038/nature11769 2013Natur.493..191C Google Scholar
Casassus, S., et al. 2013c, A&A, 553, A64 10.1051/0004-6361/201219644 2013A&A...553A..64C Google Scholar
Casassus, S., et al. 2015d, ApJ, 812, 126 10.1088/0004-637X/812/2/126 2015ApJ...812..126C CrossRefGoogle Scholar
Christiaens, V., Casassus, S., Perez, S., van der Plas, G., & Ménard, F. 2014, ApJ, 785, L12 10.1088/2041-8205/785/1/L12 2014ApJ...785L..12C Google Scholar
Cieza, L. A., et al. 2013, ApJ, 762, L12 10.1088/2041-8205/762/1/L12 CrossRefGoogle Scholar
Cieza, L. A., Schreiber, M. R., Romero, G. A., Williams, J. P., Rebassa-Mansergas, A., & Merín, B. 2012, ApJ, 750, 157 10.1088/0004-637X/750/2/157 2012ApJ...750..157C Google Scholar
Cieza, L., Williams, J., Kourkchi, E., Andrews, S., Casassus, S., Graves, S., & Schreiber, M. R. 2015, MNRAS, 454, 1909 10.1093/mnras/stv2044 2015MNRAS.454.1909C Google Scholar
Clampin, M., et al. 2003, AJ, 126, 385 10.1086/375460 2003AJ....126..385C Google Scholar
Close, L. M., et al. 2014, ApJ, 781, L30 10.1088/2041-8205/781/2/L30 2014ApJ...781L..30C Google Scholar
Corder, S., Eisner, J., & Sargent, A. 2005, ApJ, 622, L133 10.1086/429540 2005ApJ...622L.133C Google Scholar
Crida, A., Morbidelli, A., & Masset, F. 2006, Icarus, 181, 587 10.1016/j.icarus.2005.10.007 2006Icar..181..587C Google Scholar
Dent, W. R. F., Greaves, J. S., & Coulson, I. M. 2005, MNRAS, 359, 663 10.1111/j.1365-2966.2005.08938.x 2005MNRAS.359..663D Google Scholar
Dipierro, G., Price, D., Laibe, G., Hirsh, K., Cerioli, A., & Lodato, G. 2015, MNRAS, 453, L73 10.1093/mnrasl/slv105 2015MNRAS.453L..73D Google Scholar
Doğan, S., Nixon, C., King, A., & Price, D. J. 2015, MNRAS, 449, 1251 10.1093/mnras/stv347 2015MNRAS.449.1251D Google Scholar
Dong, R., Hall, C., Rice, K., & Chiang, E. 2015b, ApJ, 812, L32 10.1088/2041-8205/812/2/L32 2015ApJ...812L..32D Google Scholar
Dong, R., Zhu, Z., Fung, J., Rafikov, R., Chiang, E., & Wagner, K. 2016, ApJ, 816, L12 10.3847/2041-8205/816/1/L12 2016ApJ...816L..12D Google Scholar
Dong, R., Zhu, Z., Rafikov, R. R., & Stone, J. M. 2015a, ApJ, 809, L5 10.1088/2041-8205/809/1/L5 2015ApJ...809L...5D Google Scholar
Dunham, M. M., et al. 2015, ApJS, 220, 11 10.1088/0067-0049/220/1/11 2015ApJS..220...11D Google Scholar
Dutrey, A., et al. 2014, Nature, 514, 600 10.1038/nature13822 2014Natur.514..600D CrossRefGoogle Scholar
Eisner, J. A., Lane, B. F., Hillenbrand, L. A., Akeson, R. L., & Sargent, A. I. 2004, ApJ, 613, 1049 10.1086/423314 2004ApJ...613.1049E Google Scholar
Esau, C. F., Harries, T. J., & Bouvier, J. 2014, MNRAS, 443, 1022 10.1093/mnras/stu1211 2014MNRAS.443.1022E Google Scholar
Espaillat, C., Calvet, N., D’Alessio, P., Hernández, J., Qi, C., Hartmann, L., Furlan, E., & Watson, D. M. 2007, ApJ, 670, L135 10.1086/524360 2007ApJ...670L.135E Google Scholar
Facchini, S., Lodato, G., & Price, D. J. 2013, MNRAS, 433, 2142 10.1093/mnras/stt877 2013MNRAS.433.2142F Google Scholar
Facchini, S., Ricci, L., & Lodato, G. 2014, MNRAS, 442, 3700 10.1093/mnras/stu1149 2014MNRAS.442.3700F CrossRefGoogle Scholar
Fedele, D., et al. 2008, A&A, 491, 809 10.1051/0004-6361:200810126 2008A&A...491..809F Google Scholar
Fouchet, L., Gonzalez, J.-F., & Maddison, S. T. 2010, A&A, 518, A16 10.1051/0004-6361/200913778 2010A&A...518A..16F Google Scholar
Fukagawa, M., Tamura, M., Itoh, Y., Kudo, T., Imaeda, Y., Oasa, Y., Hayashi, S. S., & Hayashi, M. 2006, ApJ, 636, L153 10.1086/500128 2006ApJ...636L.153F Google Scholar
Fukagawa, M., et al. 2004, ApJ, 605, L53 10.1086/420699 2004ApJ...605L..53F Google Scholar
Fukagawa, M., et al. 2013, PASJ, 65, L14 10.1093/pasj/65.6.L14 2013PASJ...65L..14F Google Scholar
Garcia Lopez, R., Natta, A., Testi, L., & Habart, E. 2006, A&A, 459, 837 10.1051/0004-6361:20065575 2006A&A...459..837G Google Scholar
Garufi, A., et al. 2013, A&A, 560, A105 10.1051/0004-6361/201322429 2013A&A...560A.105G Google Scholar
Garufi, A., et al. 2016, A&A, accepted 2016A&A...accepted http://arxiv.org/abs/1601.04983 Google Scholar
Golimowski, D. A., et al. 2006, AJ, 131, 3109 10.1086/503801 2006AJ....131.3109G Google Scholar
Gonzalez, J.-F., Laibe, G., Maddison, S. T., Pinte, C., & Ménard, F. 2015, MNRAS, 454, L36 10.1093/mnrasl/slv120 2015MNRAS.454L..36G Google Scholar
Grady, C. A., Woodgate, B., Bruhweiler, F. C., Boggess, A., Plait, P., Lindler, D. J., Clampin, M., & Kalas, P. 1999, ApJ, 523, L151 10.1086/312270 1999ApJ...523L.151G Google Scholar
Grady, C. A., et al. 2001, AJ, 122, 3396 10.1086/324447 2001AJ....122.3396G Google Scholar
Grady, C. A., et al. 2009, ApJ, 699, 1822 10.1088/0004-637X/699/2/1822 2009ApJ...699.1822G Google Scholar
Grady, C. A., et al. 2013, ApJ, 762, 48 10.1088/0004-637X/762/1/48 2013ApJ...762...48G Google Scholar
Hashimoto, J., et al. 2011, ApJ, 729, L17 10.1088/2041-8205/729/2/L17 2011ApJ...729L..17H Google Scholar
Honda, M., et al. 2015, ApJ, 804, 143 10.1088/0004-637X/804/2/143 2015ApJ...804..143H Google Scholar
Hughes, A. M., et al. 2009, ApJ, 698, 131 10.1088/0004-637X/698/1/131 2009ApJ...698..131H Google Scholar
Ingleby, L., Espaillat, C., Calvet, N., Sitko, M., Russell, R., & Champney, E. 2015, ApJ, 805, 149 10.1088/0004-637X/805/2/149 2015ApJ...805..149I Google Scholar
Isella, A., Natta, A., Wilner, D., Carpenter, J. M., & Testi, L. 2010, ApJ, 725, 1735 10.1088/0004-637X/725/2/1735 2010ApJ...725.1735I Google Scholar
Isella, A., Pérez, L. M., Carpenter, J. M., Ricci, L., Andrews, S., & Rosenfeld, K. 2013, ApJ, 775, 30 10.1088/0004-637X/775/1/30 2013ApJ...775...30I CrossRefGoogle Scholar
Isella, A., Tatulli, E., Natta, A., & Testi, L. 2008, A&A, 483, L13 10.1051/0004-6361:200809641 2008A&A...483L..13I Google Scholar
Jørgensen, J. K., Visser, R., Williams, J. P., & Bergin, E. A. 2015, A&A, 579, A23 10.1051/0004-6361/201425317 2015A&A...579A..23J Google Scholar
Juhász, A., Benisty, M., Pohl, A., Dullemond, C. P., Dominik, C., & Paardekooper, S.-J. 2015, MNRAS, 451, 1147 10.1093/mnras/stv1045 2015MNRAS.451.1147J Google Scholar
Kasper, M., Apai, D., Wagner, K., & Robberto, M. 2015, ApJ, 812, L33 10.1088/2041-8205/812/2/L33 2015ApJ...812L..33K Google Scholar
Lacour, S., et al. 2015, A&A, submitted http://lib-arxiv-008.serverfarm.cornell.edu/abs/1511.09390 Google Scholar
Lin, S.-Y., Ohashi, N., Lim, J., Ho P. T. P., Fukagawa, M., & Tamura, M. 2006, ApJ, 645, 1297 10.1086/504368 2006ApJ...645.1297L Google Scholar
Looper, D. L., et al. 2010, ApJ, 714, 45 10.1088/0004-637X/714/1/45 2010ApJ...714...45L Google Scholar
Lyra, W., & Lin, M.-K. 2013, ApJ, 775, 17 10.1088/0004-637X/775/1/17 2013ApJ...775...17L Google Scholar
Maaskant, K. M., et al. 2013, A&A, 555, A64 10.1051/0004-6361/201321300 2013A&A...555A..64M Google Scholar
Manset, N., Bastien, P., Ménard, F., Bertout, C., Le van Suu, A., & Boivin, L. 2009, A&A, 499, 137 10.1051/0004-6361/200810945 2009A&A...499..137M Google Scholar
Marino, S., Casassus, S., Perez, S., Lyra, W., Roman, P. E., Avenhaus, H., Wright, C. M., & Maddison, S. T. 2015b, ApJ, 813, 76 10.1088/0004-637X/813/1/76 2015ApJ...813...76M Google Scholar
Marino, S., Perez, S., & Casassus, S. 2015a, ApJ, 798, L44 10.1088/2041-8205/798/2/L44 2015ApJ...798L..44M Google Scholar
Martin, R. G., Nixon, C., Lubow, S. H., Armitage, P. J., Price, D. J., Doğan, S., & King, A. 2014, ApJ, 792, L33 10.1088/2041-8205/792/2/L33 2014ApJ...792L..33M Google Scholar
Matter, A., et al. 2016, A&A, 586, A11 10.1051/0004-6361/201525793 http://adsabs.harvard.edu/abs/2015arXiv151003093M Google Scholar
Mayama, S., et al. 2010, Sci, 327, 306 10.1126/science.1179679 2010Sci...327..306M Google Scholar
McGinnis, P. T., et al. 2015, A&A, 577, A11 10.1051/0004-6361/201425475 2015A&A...577A..11M Google Scholar
Menu, J., van Boekel, R., Henning, T., Leinert, C., Waelkens, C., & Waters, L. B. F. M. 2015, A&A, 581, A107 10.1051/0004-6361/201525654 Google Scholar
Mendigutía, I., de Wit, W. J, Oudmaijer, R. D, Fairlamb, J. R, Carciofi, A. C, Ilee, J. D, & Vieira, R. G. 2015, MNRAS, 453, 2126 10.1093/mnras/stv1777 Google Scholar
Millar-Blanchaer, M. A., et al. 2015, ApJ, 811, 18 10.1088/0004-637X/811/1/18 2015ApJ...811...18M Google Scholar
Mittal, T., & Chiang, E. 2015, ApJ, 798, L25 10.1088/2041-8205/798/1/L25 2015ApJ...798L..25M Google Scholar
Montesinos, M., Cuadra, J., Perez, S., Baruteau, C., & Casassus, S. 2015, ApJ, 806, 253 10.1088/0004-637X/806/2/253 2015ApJ...806..253M Google Scholar
Montesinos, M., Perez, S., Casassus, S., Marino, S., Cuadra, J., & Christiaens, V. 2016, arXiv:1601.07912 2016arXiv160107912M Google Scholar
Mulders, G. D., Paardekooper, S.-J., Panić, O., Dominik, C., van Boekel, R., & Ratzka, T. 2013, A&A, 557, A68 10.1051/0004-6361/201220930 2013A&A...557A..68M Google Scholar
Muto, T., et al. 2012, ApJ, 748, L22 10.1088/2041-8205/748/2/L22 2012ApJ...748L..22M Google Scholar
Muto, T., et al. 2015, PASJ, 67, 122 10.1093/pasj/psv098 2015PASJ...67..122M Google Scholar
Najita, J., Carr, J. S., & Mathieu, R. D. 2003, ApJ, 589, 931 10.1086/374809 2003ApJ...589..931N Google Scholar
Nealon, R., Price, D. J., & Nixon, C. J. 2015, MNRAS, 448, 1526 10.1093/mnras/stv014 2015MNRAS.448.1526N Google Scholar
Nixon, C., King, A., & Price, D. 2013, MNRAS, 434, 1946 10.1093/mnras/stt1136 2013MNRAS.434.1946N Google Scholar
Ober, F., Wolf, S., Uribe, A. L., & Klahr, H. H. 2015, A&A, 579, A105 10.1051/0004-6361/201526117 2015A&A...579A.105O Google Scholar
Ohashi, N. 2008, Ap&SS, 313, 101 10.1007/s10509-007-9667-5 2008Ap%26SS.313..101O Google Scholar
Olofsson, J. et al. 2013, A&A, 552, A4 10.1051/0004-6361/201220675 2013A%26A...552A...4O Google Scholar
Owen, J. E. 2016, PASA, 33, 5 doi/10.1017/pasa.2016.2 2016pasa...33P...5 Google Scholar
Paardekooper, S.-J., & Mellema, G. 2006, A&A, 453, 1129 10.1051/0004-6361:20054449 2006A&A...453.1129P Google Scholar
Pérez, L. M., Isella, A., Carpenter, J. M., & Chandler, C. J. 2014, ApJ, 783, L13 10.1088/2041-8205/783/1/L13 2014ApJ...783L..13P Google Scholar
Perez, S., Dunhill, A., Casassus, S., Roman, P., Szulágyi, J., Flores, C., Marino, S., & Montesinos, M. 2015b, ApJ, 811, L5 10.1088/2041-8205/811/1/L5 2015ApJ...811L...5P Google Scholar
Perez, S., et al. 2015a, ApJ, 798, 85 10.1088/0004-637X/798/2/85 2015ApJ...798...85P Google Scholar
Piétu, V., Guilloteau, S., & Dutrey, A. 2005, A&A, 443, 945 10.1051/0004-6361:20042050 2005A&A...443..945P Google Scholar
Pinilla, P., Benisty, M., & Birnstiel, T. 2012, A&A, 545, A81 10.1051/0004-6361/201219315 2012A&A...545A..81P Google Scholar
Pinilla, P., et al. 2015, A&A, 584, A16 10.1051/0004-6361/201526655 2015A&A...584A..16P Google Scholar
Pinte, C., Dent, W. R. F., Ménard, F., Hales, A., Hill, T., Cortes, P., & de Gregorio-Monsalvo, I. 2016, ApJ, 816, 25 10.3847/0004-637X/816/1/25 2016ApJ...816...25P Google Scholar
Pohl, A., Pinilla, P., Benisty, M., Ataiee, S., Juhász, A., Dullemond, C. P., Van Boekel, R., & Henning, T. 2015, MNRAS, 453, 1768 10.1093/mnras/stv1746 2015MNRAS.453.1768P Google Scholar
Pontoppidan, K. M., Blake, G. A., & Smette, A. 2011, ApJ, 733, 84 10.1088/0004-637X/733/2/84 2011ApJ...733...84P Google Scholar
Pontoppidan, K. M., Blake, G. A., van Dishoeck, E. F., Smette, A., Ireland, M. J., & Brown, J. 2008, ApJ, 684, 1323 10.1086/590400 2008ApJ...684.1323P Google Scholar
Quanz, S. P., Amara, A., Meyer, M. R., Girard, J. H., Kenworthy M. A., & Kasper, M. 2015, ApJ, 807, 64 10.1088/0004-637X/807/1/64 2015ApJ...807...64Q Google Scholar
Quanz, S. P., Amara, A., Meyer, M. R., Kenworthy, M. A., Kasper, M., & Girard, J. H. 2013, ApJ, 766, L1 10.1088/2041-8205/766/1/L1 2013ApJ...766L...1Q Google Scholar
Quillen, A. C. 2006, ApJ, 640, 1078 10.1086/500165 2006ApJ...640.1078Q Google Scholar
Rafikov, R. R. 2016, ApJ, submitted http://arxiv.org/abs/1601.03009 Google Scholar
Rameau, J., Chauvin, G., Lagrange, A.-M., Thébault, P., Milli, J., Girard, J. H., & Bonnefoy, M. 2012, A&A, 546, A24 10.1051/0004-6361/201219736 2012A&A...546A..24R Google Scholar
Regály, Z., Juhász, A., Sándor, Z., & Dullemond, C. P. 2012, MNRAS, 419, 1701 10.1111/j.1365-2966.2011.19834.x 2012MNRAS.419.1701R CrossRefGoogle Scholar
Reggiani, M., et al. 2014, ApJ, 792, L23 10.1088/2041-8205/792/1/L23 2014ApJ...792L..23R Google Scholar
Roberge, A., Weinberger, A. J., & Malumuth, E. M. 2005, ApJ, 622, 1171 10.1086/427974 2005ApJ...622.1171R Google Scholar
Rodigas, T. J., Follette, K. B., Weinberger, A., Close, L., & Hines D. C. 2014, ApJ, 791, L37 10.1088/2041-8205/791/2/L37 2014ApJ...791L..37R Google Scholar
Rosenfeld, K. A., Chiang, E., & Andrews, S. M. 2014, ApJ, 782, 62 10.1088/0004-637X/782/2/62 2014ApJ...782...62R Google Scholar
Rosenfeld, K. A., et al. 2012, ApJ, 757, 129 10.1088/0004-637X/757/2/129 2012ApJ...757..129R Google Scholar
Sallum, S., et al. 2015, Nature, 527, 342 10.1038/nature15761 2015Natur.527..342S Google Scholar
Salyk, C., Blake, G. A., Boogert, A. C. A., & Brown, J. M. 2009, ApJ, 699, 330 10.1088/0004-637X/699/1/330 2009ApJ...699..330S Google Scholar
Salyk, C., Pontoppidan, K., Corder, S., Muñoz, D., Zhang, K., & Blake, G. A. 2014, ApJ, 792, 68 10.1088/0004-637X/792/1/68 2014ApJ...792...68S Google Scholar
Schisano, E., Covino, E., Alcalá, J. M., Esposito, M., Gandolfi D., & Guenther, E. W. 2009, A&A, 501, 1013 10.1051/0004-6361/200811073 2009A&A...501.1013S Google Scholar
Schneider, P. C., France, K., Günther, H. M., Herczeg, G., Robrade, J., Bouvier, J., McJunkin, M., & Schmitt, J. H. M. M. 2015, A&A, 584, A51 10.1051/0004-6361/201425583 2015A&A...584A..51S Google Scholar
Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARAA, 25, 23 10.1146/annurev.aa.25.090187.000323 1987ARA&A..25...23S CrossRefGoogle Scholar
Tang, Y.-W., Guilloteau, S., Piétu, V., Dutrey, A., Ohashi, N., & Ho, P. T. P. 2012, A&A, 547, A84 10.1051/0004-6361/201219414 2012A&A...547A..84T Google Scholar
Tatulli, E., et al. 2011, A&A, 531, A1 10.1051/0004-6361/201016165 2011A&A...531A...1T Google Scholar
Testi, L., et al. 2014, Protostars and Planets VI, 10.2458/azu_uapress_9780816531240-ch015 2014prpl.conf..339T339–361Google Scholar
Testi, L., et al. 2015, ApJ, 812, L38 10.1088/2041-8205/812/2/L38 2015ApJ...812L..38T Google Scholar
Thi, W.-F., et al. 2014, A&A, 561, A50 10.1051/0004-6361/201322150 2014A&A...561A..50T Google Scholar
van Boekel, R., et al. 2004, Nature, 432, 479 10.1038/nature03088 2004Natur.432..479V Google Scholar
van Dishoeck, E. F., van der Marel, N., Bruderer, S., & Pinilla, P. 2016, MNRAS, accepted http://arxiv.org/abs/1511.03497 Google Scholar
van der Marel, N., et al. 2013, Science, 340, 1199 2013Sci...340.1199V Google Scholar
van der Marel, N., Pinilla, P., Tobin, J., van Kempen, T., Andrews S., Ricci, L., & Birnstiel, T. 2015b, ApJ, 810, L7 10.1088/2041-8205/810/1/L7 2015ApJ...810L...7V Google Scholar
van der Marel, N., van Dishoeck, E. F., Bruderer, S., Andrews, S. M., Pontoppidan, K. M., Herczeg, G. J., van Kempen, T., & Miotello, A. 2016, A&A, 585, A58 10.1051/0004-6361/201526988 2016A&A...585A..58V Google Scholar
van der Marel, N., van Dishoeck, E. F., Bruderer, S., Pérez, L., & Isella, A. 2015a, A&A, 579, A106 10.1051/0004-6361/201525658 2015A&A...579A.106V Google Scholar
van der Plas, G., Casassus, S., Ménard, F., Perez, S., Thi, W. F., Pinte, C., & Christiaens, V. 2014, ApJ, 10.1088/2041-8205/792/2/L25 2014ApJ...792L..25V 792, L25 Google Scholar
van der Plas, G., van den Ancker, M. E., Acke, B., Carmona, A., Dominik, C., Fedele, D., & Waters, L. B. F. M. 2009, A&A, 500, 1137 10.1051/0004-6361/200811148 2009A&A...500.1137V Google Scholar
van der Plas, G., van den Ancker, M. E., Fedele, D., Acke, B., Dominik, C., Waters, L. B. F. M., & Bouwman, J. 2008, A&A, 485, 487 10.1051/0004-6361:20078867 2008A&A...485..487V Google Scholar
van der Plas, G., van den Ancker, M. E., Waters, L. B. F. M., & Dominik C. 2015, A&A, 574, A75 10.1051/0004-6361/201425052 2015A&A...574A..75V Google Scholar
Wagner, K., Apai, D., Kasper, M., & Robberto, M. 2015, ApJ, 813, L2 10.1088/2041-8205/813/1/L2 2015ApJ...813L...2W Google Scholar
Wahhaj, Z., et al. 2015, A&A, 581, A24 10.1051/0004-6361/201525837 2015A&A...581A..24W Google Scholar
Wang, J. J., et al. 2015, ApJ, 811, L19 10.1088/2041-8205/811/2/L19 2015ApJ...811L..19W Google Scholar
Weidenschilling, S. J. 1977, MNRAS, 180, 57 10.1093/mnras/180.1.57 1977MNRAS.180...57W Google Scholar
Whitney, B. A., Robitaille, T. P., Bjorkman, J. E., Dong, R., Wolff M. J., Wood, K., & Honor, J. 2013, ApJS, 207, 30 10.1088/0067-0049/207/2/30 2013ApJS..207...30W Google Scholar
Williams, J. P., & Cieza, L. A. 2011, ARAA, 49, 67 10.1146/annurev-astro-081710-102548 2011ARA&A..49...67W Google Scholar
Zhang, K., Isella, A., Carpenter, J. M., & Blake, G. A. 2014, ApJ, 791, 42 10.1088/0004-637X/791/1/42 2014ApJ...791...42Z Google Scholar
Zhu, Z., & Baruteau, C. 2015, preprint 2015arXiv151103497Z (arXiv 1511.03497)Google Scholar
Zhu, Z., & Stone, J. M. 2014, ApJ, 795, 53 10.1088/0004-637X/795/1/53 2014ApJ...795...53Z Google Scholar
Zuckerman, B. 2001, ARAA, 39, 549 10.1146/annurev.astro.39.1.549 2001ARA Google Scholar
Figure 0

Figure 1. Figure adapted from Whitney et al. (2013, ©AAS, reproduced by permission), with a three-dimensional radiative transfer prediction in JHK for a warped disk configuration. The field is 4 AU on a side.

Figure 1

Figure 2. Comparison between the observed H-band polarised intensity image of HD 142527 (a- from Avenhaus et al. 2014a) and three-dimensional radiative transfer predictions (b- from Marino et al. 2015a, this is an updated version of their Figure 2). The kinematics of C18O(2 − 1) emission (Perez et al. 2015a) give the orientation of the outer disk; the white contours in a- correspond to systemic velocities, so that the PA of the outer disk lies at ~ 160° East of North, as indicated on b-. The inner disk shadows cast on the outer disk are best reproduced with a PA of 172°, the curvature of their outline (or silhouette) is reminiscent of the observations for the Eastern side. The similarities with the observations are particularly good given the idealisations of the model, which assumes a circular cavity.

Figure 2

Figure 3. Summary of Cycle 0 band 7 observations, from MEM maps, with continuum in red, HCO+(4 − 3) in green, and CO(3 − 2) in blue (adapted from Casassus et al. 2013b). x − and y − show offset from the star, in arcsec. Velocities have been restricted to highlight the fainter structures seen in HCO+, which are otherwise dwarfed by the fast HCO+ central emission.

Figure 3

Figure 4. Figure adapted from Casassus et al. (2015a), showing the observed CO(6 − 5) kinematics in the central regions of HD 142527, and comparing with model predictions for the accretion through the warp represented in Figure 5 (convolved at the resolution of the centroid map). The origin of coordinates is set to the stellar position.

Figure 4

Figure 5. Sketch of the accretion kinematics through the warp in HD 142527 (from Casassus et al. 2015a). The crescent in hues of red and purple represents the distribution of mm-sized grains. Inside the warp, as it connects the two disk inclination material flows along the dashed and curved arrows.

Figure 5

Figure 6. Figure adapted from Casassus et al. (2015d). Upper row: multi-frequency data of HD 142527 brought to a common uv-coverage. The ALMA data have been filtered for the ATCA response with Monte Carlo simulations of ATCA observations on deconvolved models of the ALMA data. (a): restored ATCA image at 34 GHz. (b): average of Monte Carlo (MC) simulations of ATCA observations on the ALMA band 7 data. (c): MC simulations of ATCA observations on the ALMA band 9 data. Lower row: Grey-body diagnostics inferred from the multi-frequency data. (a): optical depth map at the reference frequency of 345 GHz. (b): line of sight emissivity index map β, with ATCA specific intensity contours in red. (c): root-mean-square uncertainties on the emissivity index map. (d): line of sight temperature, Ts.

Figure 6

Figure 7. Figure adapted from van der Marel et al. (2015b, their Figure 1, ©AAS, reproduced with permission). VLA observations of IRS 48 at optically thin frequencies confirm the extreme lopsidedness of this TD. VLA observations at 34 GHz are shown in red contours—the tail towards the NE is likely due to stellar emission. ALMA band 9 observations at 680 GHz are shown in black contours, after filtering for the VLA response. The field is 2 arcsec on a side. While at 34 GHz the crescent appears to be somewhat more compact, the role of optical depth effects in widening the 680 GHz signal remain to be quantified.

Figure 7

Figure 8. Figure adapted from Marino et al. (2015b, their Figure 3). VLA observations of MWC 758 compared to a deconvolved model of the ALMA band 7 visibilities. The VLA signal is shown in red contours—the red ellipse corresponds to the synthesised beam. The blue scale corresponds to an ‘MEM’ non-parametric model of the 337 GHz ALMA data—the blue ellipse corresponds to an elliptical Gaussian fit to the point-spread-function of the ‘MEM’ algorithm.

Figure 8

Figure 9. Left: Spiral model involving a planetary mass companion, from Dong et al. (2015a, part of their Fig. 4, ©AAS, reproduced with permission). The companion is highlighted as a green dot—and the field is rotated so that the companion approximately matches the location of the VLA clump seen in MWC 758. Right (adapted from Marino et al. 2015b): VLA data, in red contours, and ALMA 345GHz, in blue contours, overlaid on the SPHERE Y-band polarised intensity image (Benisty et al. 2015), in grey scale.

Figure 9

Figure 10. Gallery of optical/IR spirals. From left to right, we show scattered-light images of AB Aur (Fukagawa et al. 2004, part of their Fig. 3, ©AAS, reproduced with permission), HD 141569A (Clampin et al. 2003, part of their Fig. 8, ©AAS, reproduced with permission), HD 135344B (Garufi et al. 2013, part of their Fig. 1, ©ESO, reproduced with permission), and HD 100453 (Wagner et al. 2015, part of their Fig. 2, ©AAS, reproduced with permission). All figures have been reproduced by permission of the AAS or A&A. In HD 100453, we have highlighted two intensity dips where the two-armed spiral pattern seem to stem from. These dips are very reminiscent of the HD 142527 shadows, which also seem to be at the root of spirals.

Figure 10

Figure 11. Radio spirals from HD 142527 (Figure adapted from Christiaens et al. 2014). The CO(2 − 1) peak signal is shown on the left, while CO(3 − 2) is shown on the right (the gray ellipse correspond to the clean beams). x − and y − axis indicate angular offset from the star, along RA and Dec., and in arcsec. Since the distance to HD 142527 is about 140 pc, the spirals are seen to extend out to 700 pc in radius. The lozenge symbols indicate the most conspicuous IR spiral, from Fukagawa et al. (2006). The inverted-V decrement in peak CO intensity, seen in both transitions, is due to interstellar absorption at velocities that correspond to these locations in the Keplerian outer disk.

Figure 11

Figure 12. Spirals launched from the temperature forcing of the outer disk by shadows projected from a tilted inner disk—as in HD 142527 (and perhaps also in HD 100453). The image shows a snapshot of the density field at 250 orbits, under the effect of shadows aligned in the East–West direction (from Montesinos et al. 2016).